Sie sind auf Seite 1von 13

Published October 19, 2017

Review & Analysis–Soil Physics & Hydrology

Numerical Solution of Richards’ Equation:


A Review of Advances and Challenges
The flow of water in partially saturated porous media is of importance in fields
Matthew W. Farthing
Coastal and Hydraulics Lab. such as hydrology, agriculture, environment and waste management. It is also
US Army Engineer Research and one of the most complex flows in nature. The Richards’ equation describes the
Development Center flow of water in an unsaturated porous medium due to the actions of grav-
3909 Halls Ferry Rd ity and capillarity neglecting the flow of the non-wetting phase, usually air.
Vicksburg, MS 39180 Analytical solutions of Richards’ equation exist only for simplified cases, so
most practical situations require a numerical solution in one- two- or three-
Fred L. Ogden* dimensions, depending on the problem and complexity of the flow situation.
Dep. of Civil and Despite the fact that the first reasonably complete conservative numerical solu-
Architectural Engineering tion method was published in the early 1990s, the numerical solution of the
Univ. of Wyoming Richards’ equation remains computationally expensive and in certain circum-
1000 E. Univ. Ave. Dep. 3295
stances, unreliable. A universally robust and accurate solution methodology has
Laramie, WY 82071
not yet been identified that is applicable across the range of soils, initial and
boundary conditions found in practice. Existing solution codes have been modi-
fied over years to attempt to increase robustness. Despite theoretical results on
the existence of solutions given sufficiently regular data and constitutive rela-
tions, our numerical methods often fail to demonstrate reliable convergence
behavior in practice, especially for higher-order methods. Because of robust-
ness, the lack of higher-order accuracy and computational expense, alternative
solution approaches or methods are needed. There is also a need for better
documentation of improved solution methodologies and benchmark test prob-
lems to facilitate consistent advances and avoid re-inventing of the wheel.

T
his review paper is intended to serve as a touchstone on the state of the
science of calculating the flow of water through unsaturated porous me-
dia. As active researchers we believe it is good to occasionally take stock
in what has been accomplished up to date, where things may be headed, and what
important issues remain. This review concentrates on computational modeling of
flow through the unsaturated zone via Richards’ equation.
This effort can help provide some focus to the research community and serve
to help propel new research forward, particularly by those just joining the field.
There are many practical problems that require calculation of one-, two-, and three-
dimensional fluxes in the unsaturated zone. For a much broader review of modeling
soil processes, we recommend the recent paper from Vereecken et al. (2016). A
detailed review of mathematical models of infiltration can be found (Assouline,
2013), while Paniconi and Putti (2015) provide historical perspective and an
overview of modeling Richards’ equation in the context of catchment hydrology.
Core Ideas The equation attributed to Richards (1931) that describes the flow of water
• The numerical solution of Richards’ through unsaturated porous media under the action of capillarity and gravity was
equation remains challenging. first published by the English mathematician and physicist Lewis Fry Richardson
• Space/time discretization affects both in 1922 (Richardson, 1922). Therefore, the equation would rightly be called
computational effort and accuracy. “Richardsons’ equation”, although in this paper we continue to call it by its
common name, Richards’ equation.
• Adaption of space and time
discretizations produces benefits.
Soil Sci. Soc. Am. J. 81:1257–1269
• Dissemination of codes and improved
doi:10.2136/sssaj2017.02.0058
documentation are needed. Received 16 Feb. 2017.
Accepted 4 June 2017.
• Recent reformulation of one- *Corresponding author: (fogden@uwyo.edu)
dimensional Richards’ equation © Soil Science Society of America. This is an open access article distributed under the CC BY-NC-ND
shows promise. license (http://creativecommons.org/licenses/by-nc-nd/4.0/)


Soil Science Society of America Journal
Richards’ equation can be seen as a simplification of the ∂y ∂   ∂y  
standard two-phase flow formulation for a gas and water phase in  S s S a ( y ) + c ( y )  ∂ t − ∂ z  K ( y )  ∂ z −1   =
0 [5]
  
a porous medium where the pressure gradient required to drive
flow of the gas phase is ignored due to the large mobility contrast where Ss [L–1] is the specific storage and Sa [-] is the saturation of
between the water and gas phases. One of the most interesting the aqueous phase (Miller et al., 1998). While ¶q/¶t = c(y)¶y/¶t
aspects of the Richards’ equation is that despite its ease of holds at the continuous level through the chain rule, the
derivation, it is arguably one of the most difficult equations to nonlinearity of c(y) makes it impossible to maintain this equality
reliably and accurately solve in all of hydrosciences. uniformly at the discrete level, and the pure head-based form in Eq.
Richards’ equation can be written in a number of forms, [4] or [5] is not generally mass conservative. There are techniques
namely the water content, mixed water content and capillary like a chord-slope approximation (Rathfelder and Abriola, 1994)
head form, and the head form. In one-dimension, the so-called that can be used in specific cases or flux updating that can promote,
“mixed water content form”, because it mixes the water content θ but not ensure, mass conservation (Kirkland et al., 1992). In
with the capillary head y(θ), is: general though, high accuracy in the time integration of Eq. [4] is
essential to avoid mass balance errors (Tocci et al., 1997), and Eq.
∂q ∂   ∂ y(q)  
=  K (q )  −1    [1] [3] together with expansion of the nonlinearity θ(y) in the storage
∂t ∂z   ∂z 
change term, ¶θ/¶t, (Allen and Murphy, 1985; Celia et al., 1990)
where z is vertical coordinate (positive downward)[L]; t is time is preferred when low-order time integration is used.
[T]; q equals q(z,t) equals volumetric soil moisture content [-]; Richards’ equation is a nonlinear, degenerate elliptic-
y(q) is empirical soil hydraulic capillary head function [L]; K(q) parabolic partial differential equation (List and Radu, 2016).
= empirical unsaturated hydraulic conductivity function [L T–1]. The head-based forms of the equation, Eq. [3–5], transition
Equation [1] is often written solely as a function of the from parabolic to elliptic as the solution domain nears saturation
water content by introducing the chain rule: unless storage effects are included (Ss > 0), in which case Eq. [5]
remains parabolic. The nonlinearity and transition of type in
∂q ∂  ∂q 
=  D(q) − K (q)  [2] different portions of the problem domain make it very difficult to
∂t ∂z  ∂z  solve using traditional analytical techniques, and it is impossible
where D(θ) = K(θ)(¶y(θ)/¶θ) is referred to as the soil-water to solve in closed form except for a small number of special cases
diffusivity [L2 T–1]. As in Eq. [1], the first term in parentheses (Miller et al., 1998). This difficulty is only exacerbated by the
captures the effects of capillarity, while second term in range of initial and boundary conditions encountered in practice
parentheses represents the effect of gravity-driven flux. as well as by the nature of common soil water constitutive relations
In uniform soils the water content or mixed water content that can lead to solutions with low regularity (Alt and Luckhaus,
forms of Richards’ equation are valuable because water content 1983). Specifically, the solution depends on two empirical,
is a continuous variable. However, in nature soils are seldom highly nonlinear soil water constitutive relations: the unsaturated
uniform over significant length scales, and layered soils are hydraulic conductivity function K(θ) that can be constant or
ubiquitous. In layered soils the water content is discontinuous very near zero for non-positive values of the capillary head, and
across layer interfaces because of unique unsaturated capillary the capillary head function y(θ) that can take on arbitrary small
head relations in the different soil layers (Assouline, 2013). values for relative saturations near 100%. These extremes lead
Rather, the capillary head (y) is continuous, and it is better to to degeneracy in the solution of Richards’ equation. The soil
write the Richards’ equation with capillary head as the dependent constitutive relations may not be smoothly differentiable at these
variable and evaluate the moisture content in terms of y, θ = extremes and can further have very high slopes and hysteresis,
θ(y). This can be done in either a mixed form: and be discontinuous at low relative saturations. Furthermore,
∂ q(y) ∂  infiltration into dry soils often leads to sharp wetting fronts
 ∂y  
−  K ( y)  −1   =
0 [3] that result in extremely large spatial gradients of soil hydraulic
∂t ∂z   ∂z  
properties (Zha et al., 2017). These nonlinearities and the
or via the chain rule as the fully head-based form: degeneracy make the design and analysis of numerical schemes for
the Richards’ equation very difficult (Miller et al., 2013).
∂y ∂   ∂ y   [4] Exact one-dimensional solutions of Richards’ equation have
c ( y) − K ( y)  −1   =
0
∂ t ∂ z   ∂z   been derived for a few specialized forms of the constitutive relations
where c(y) equals ¶q/¶y equals the specific moisture describing the soil water retention and the unsaturated hydraulic
capacity [L–1]; K(y) equals the hydraulic conductivity conductivity functions (Rogers et al., 1983; Broadbridge and
function written as a function of y[L T–1]. White, 1988; Sander et al., 1988; Barry and Sander, 1991; Barry et
Equations [3] and [4] can both be used to solve unsaturated al., 1993; Ross and Parlange, 1994). However, these solutions are
and saturated flow problems. One may also encounter Eq. [4] in not generally applicable because either the functional forms are
a form that includes fluid compressibility effects dissimilar from widely used constitutive relations that represent
real soils, and/or the solutions impose strict requirements on

1258 Soil Science Society of America Journal


the initial and boundary conditions. These exact solutions are Infiltration is often assumed to be a one-dimensional process
valuable for verifying numerical solutions. in the vertical direction. Or et al. (2015) demonstrated that higher
Given the lack of a general closed-form solution of the dimensional approaches are generally necessary only at scales
Richards’ equation, numerical solvers are required when accurate less than approximately 10 m. The one-dimensional solution
unsaturated flow simulation capability is desired. Of course, methodology developed by Celia et al. (1990), which uses modified
many approximate, empirical, or completely arbitrary methods Picard iterations to improve mass conservation, has become the
have been developed over the years to simulate the flow of standard numerical approach. It remains essentially the method that
water through unsaturated soils because of the real or perceived is used in many production codes including the USDA Hydrus-1D
difficulties posed by the numerical solution of Richards’ equation. Richards’ equation solver (Simunek et al., 2005).
Here we ignore those, and we solely consider problems where fully As a result of the one-dimensional vertical assumption, large-
dynamical simulations in unsaturated porous media are required. scale models are able to replace a fully coupled, three-dimensional
Note that in the following we focus on head-based forms of solution with many, independent one-dimensional calculations.
Richards’ equation that are valid for unsaturated or fully saturated Because of reliability issues however, current large-scale models of
(i.e., variably saturated) conditions. When necessary, we will use land/atmosphere interaction often use approximate or conceptual
the term “strictly unsaturated” to refer to conditions or techniques methods to model vadose zone fluxes. Those approximations
that are restricted to relative saturation values less than 100%. usually require assumption of runoff generation mechanism,
The complexities of solving the Richards’ equation were and are used in land-surface schemes such as: VIC (Liang et al.,
discussed by Gray and Hassanizadeh (1991), who referred to the 1994), JULES (Cox et al., 1999), NOAH (Ek et al., 2003), CLM
formulation as “paradoxical and overly simplistic”. Since then, (Niu et al., 2007), NOAH-MP (Yang et al., 2011). The reality
many authors have presented critiques of Richards’ equation is that the numerical solution of Richards’ equation is avoided
and existing solution techniques. A far-from-exhaustive sample in these schemes because of concerns regarding computational
includes (Forsyth et al., 1995; Tocci et al., 1997; van Dam and expense and solution reliability, as well as questions regarding
Feddes, 2000; Vogel et al., 2001; Kavetski et al., 2001a; Farthing the appropriateness of applying the equation to represent the
et al., 2003b; Bause and Knabner, 2004; Manzini and Ferraris, behavior of large model elements or grids.
2004; D’Haese et al., 2007; Kees et al., 2008; Vogel and Ippisch, The penalty in using a Richards’ solver in hydrological
2008; Juncu et al., 2011; Berninger et al., 2014; Lipnikov et al., simulation with thousands or millions of one-dimensional solutions
2016). Our intention here is not to repeat what those authors comes from the fact that while many of these may converge with
have written, rather it is to assess the state of the art, limitations, rapidity and ease, there are conditions where some of them will
needs, and alternatives. not (Niswonger and Prudic, 2009) for a variety of reasons. In those
cases, where convergence is slow or not possible, the code must be
Calculation of Vadose Zone Fluxes able to identify this condition and move to the next time step with
In the Richards’ equation, the physical properties of some approximate solution. This problem affects the reliability of
the medium are represented as nonlinear coefficients. The hydrological simulators that use Richards’ equation. The lack of
nonlinearities introduce a number of challenges as alluded solution reliability imposes a large penalty on the use of Richards’
to in the introduction. For example, with the widely used van equation solvers when convergence is unattainable. This penalty
Genuchten and Mualem constitutive relations, the capillary takes the form of computational expense, slowdown of parallel
head function, y (θ), and the specific moisture capacity, ¶θ/¶y, algorithms waiting for detection and approximations in the case
approach zero as the moisture content in the porous medium of non-convergent solutions, and solutions with large errors,
nears saturation, while the soil-water diffusivity, D(θ), can grow including significant mass-balance errors.
arbitrarily large (van Genuchten, 1980). Loosely, this behavior The convergence abilities and computational efficiency (or
leads to a condition known as degeneracy, when the solution of lack thereof ) for a particular Richards’ equation solver depend
Richards’ equation is hampered by coefficients that take values on many factors. The quality of the software engineering and
near zero or infinity. Classical analytical results on the existence, implementation of the numerics are obviously critical. The
uniqueness, and regularity of weak solutions can be found essential numerical components include the spatial and temporal
in (Van Duyn and Peletier, 1982; Alt and Luckhaus, 1983), approximations used to transition from an initial boundary value
while Gilding (1991) provides analysis of infiltration in strictly problem posed at the continuous level to a discrete system, as well
unsaturated domains, including the positivity and boundedness as the nonlinear and linear algebra techniques that are used to
of soil moisture and conditions for finite speed of propagation solve the resulting discrete systems and produce an approximate
of wetting fronts. In practice, the difficulties driven by these solution (Trangenstein, 2013). Here, we focus on the numerical
nonlinearities often manifest as rapid changes in capillary head discretization techniques themselves and not the very important
and soil moisture around infiltration fronts as well as complex software engineering issues surrounding the development of high-
material coefficients that are potentially non-continuously quality scientific codes (McConnell, 2004; Bartlett, 2009; Heroux
differentiable near full saturation (Miller et al., 1998). and Willenbring, 2009). We summarize briefly the current state of

www.soils.org/publications/sssaj 1259
the art and identify outstanding issues for the primary numerical an example, Fig. 2 illustrates a typical piecewise linear Lagrange
components of a Richards’ equation solver below. polynomial, wI, that is the building block for most finite element
models for Richards’ equation. It takes a value of 1 at vertex,
Spatial Discretization I, and is zero at the other vertices of the triangulation, Mh. Of
The overwhelming majority of Richards’ equation solvers course, Eq. [6] and [7] are only intended to provide some context
employ either a finite difference, finite volume, or finite element for the following discussion, and we have neglected any details of
approximation in space. Hybrid combinations are also possible the solution approximation or evaluation of the integrals in Eq.
(Helmig, 2011). Table 1 provides a representative list of research [6] and [7] (Kees et al., 2008).
and production codes along with their spatial approximations. The spatial accuracy of all of these approximation methods
The details of these approaches vary. They all build on a discrete depends on several factors. In particular, it is generally a function
partition or tessellation of the spatial domain of interest. The of the local mesh geometry, which can be expressed through a
tessellation may be a structured grid or more generally an characteristic length scale and a measure of the aspect ratio or
unstructured mesh. Figure 1 illustrates a simple unstructured distortion (Trangenstein, 2013). Loosely, the formal accuracy of
mesh of a polygonal spatial domain, W, consisting of a union of a convergent scheme can be related to the characteristic length
non-empty, non-overlapping triangles, Mh = {We}. scale of the mesh raised to some power, which we will refer to as
Each scheme must then approximate the spatial variation the scheme’s order. There is a traditional rule of thumb that large
of material properties and solution variables on this mesh and angles (hence aspect ratios) negatively impact approximation
enforce the conservation statement and solution dynamics accuracy (Shewchuk, 2002). On the other hand, it is also possible
expressed by Richards’ equation in some way. For example, through mesh optimization techniques to exploit solution
conservation might be enforced point-wise in a traditional finite behavior and gain higher accuracy through the appropriate use of
difference scheme. Or, it might be enforced in an integral sense element anisotropy (Pain et al., 2001; Mostaghimi et al., 2015).
using test functions that are constant over a control volume, V, in The spatial discretization size is relevant beyond just the
a finite volume method formal accuracy. Convergence of a method relates to how well
the scheme approximates a solution to Richards’ equation. It
∂q
∫ d x − ∫ ∇⋅[ K (y)(∇ y −1z ) ] d x = 0 [6] does not address whether or not Richards’ equation is a valid
V ∂t V
approximation for unsaturated flow at the length scales used
Here, 1z is a vector denoting the vertical axis. V could correspond in the mesh, however. For example, it is unlikely that a three-
to a mesh element, We, in “cell-centered” scheme or a dual control dimensional solution of Richards’ equation on a large, multiple-
volume, W  , in a “vertex-centered” scheme (Fig. 1) (Huber and kilometer-scale voxel can accurately represent nature or be
I
Helmig, 2000; Manzini and Ferraris, 2004). Finite element physically meaningful because it will violate the representative
methods are also weighted residual methods that enforce elementary volume assumption (Or et al., 2015).
Richards’ equation weakly, For reference, Fig. 3 shows three commonly used one-
∂q dimensional discretizations. From left to right these spatial
∫ w d x − ∫ ∇⋅[ K (y)(∇ y −1z ) ] w d =
x 0, ∀ w∈W (W) [7] discretizations represent increasing degrees of complexity, and
W ∂t W
increasing degrees of accuracy. The uniform discretization is
Finite element methods impose a richer structure on the space simplest, but the nature of the solution requires fine resolution
of weighting functions, W(W) (Ern and Guermond, 2004). As everywhere. The variable discretization provides high resolution
near the land surface, where accurately simulating the change
Table 1. Alphabetical list of representative research and pro- in water content with time is most important for accurate
duction Richards’ equation codes. partitioning of rainfall or energy (Downer and Ogden, 2004).
Spatial The variable spatial discretization also allows coarsening
Code approximation Reference away from the land surface, which can reduce computational
ADH CG FEM† Howington et al. 1999
CATHY CG FEM Camporese et al., 2010
FEHM CV FEM Zyvoloski, 2007
FEFLOW CG FEM Trefry and Muffels, 2007
HydroGeoSphere CV FEM Brunner and Simmons, 2012
HYDRUS CG FEM Simunek et al., 2008
ParFlow CCD Kollet and Maxwell, 2006
RichardsFOAM FV Orgogozo et al., 2014
TOUGH IFD Finsterle et al., 2008
VS2D CCD Healy, 2008
WASH123D CG FEM Yeh et al., 2011
†C G FEM, Continuous Galerkin Finite Element Method; CV FEM,
Control Volume Finite Element Method; CCD, Cell-Centered
Differences; FV, Finite Volume; IFD, Integrated Finite Differences. Fig. 1. Example triangulation with dual mesh.

1260 Soil Science Society of America Journal


burden without significantly affecting the solution accuracy. like the one sketched on Fig. 3C. Many researchers have shown
The adaptive spatial discretization is a much more complicated that adaptive resolution can greatly improve accuracy and
situation wherein the solver places computational points that robustness (Bause and Knabner, 2004; Li et al., 2007b). Indeed
allow the most accurate simulation of the time evolution of the even one-dimensional approximations have shown that they can
wetting front. pose a significant computational burden and benefit from adaptive
The past three decades have seen the introduction and resolution, depending on the combination of soil properties, initial
maturation of a number of spatial approximation methods for conditions, and boundary forcing (Miller et al., 2006). However,
elliptic problems that are robust for high contrasts in material despite the introduction of sophisticated adaption techniques
properties and heterogeneity while providing locally conservative that have have matured in other computational mechanics fields
velocities for transport. These schemes include cell and vertex- (Schwab, 1998; Rannacher, 2001; Fidkowski and Darmofal,
centered finite volume schemes, mimetic finite differences, 2011), including so-called h adaption which modifies the local
multi-point flux approximation (MPFA) techniques, mixed mesh spacing (Pettway et al., 2010) as well as methods that
finite elements, and discontinuous Galerkin methods (Chavent vary the local mesh spacing and approximation order (i.e., h-p
and Roberts, 1991; Arbogast et al., 1995; Aavatsmark et al., adaption) (Solin and Kuraz, 2011), spatial adaption for Richards’
1998; Campbell et al., 2002; Dawson et al., 2004; Klausen and equation is still not widely used in practice.
Russell, 2004; Trangenstein, 2013).
These techniques have been largely successful for multiphase Temporal Discretization
flow problems when applied to the global flow (or pressure) As with spatial approximations, solution of time-varying
equation in fractional flow formulations (Chavent et al., 1984; problems requires partition of the temporal domain of interest.
Gerritsen and Durlofsky, 2005; Hoteit and Firoozabadi, 2005; Generating a discrete partition for an interval, [0,T] 0 = t0 < t1
Chen et al., 2006). They have also been applied to Richards’ < ... < tm = T, is much simpler than a complex three-dimensional
equation by a number of researchers (Bergamaschi et al., 1999; Li domain. Nevertheless, accuracy remains a function of the time
et al., 2007a; Arrarás et al., 2009; Klausen et al., 2008; Kumar et step size. Schemes can be characterized in terms of their local
al., 2009). However, their direct translation to variably saturated discretization error, which is the error that would originate over
flow has proven challenging. one time step, Δt(n+1) = t(n+1) –tn, starting from the true solution
To achieve comparable robustness, non-trivial modifications at tn. The error that accumulates over the entire interval [0,T] is
like low-order quadrature and mass lumping (Forsyth and known as the global error (Hairer et al., 1996). We will, somewhat
Kropinski, 1997; Arbogast et al., 1998; Woodward and Dawson, loosely, refer to a method as order p if its local error is order p+1
2000; Farthing et al., 2003b; Pop et al., 2004; Belfort et al.,
2009; Younes et al., 2013) have often been required along with ∈n+1 ≤C (D t n+1 ) p+1 [8]
significant work on methods for solving the resulting nonlinear Here C is a constant that is independent of the time step (Hairer
discrete problems (Paniconi et al., 1991; Bergamaschi et al., 1999; et al., 1996).
Bause and Knabner, 2004; List and Radu, 2016). As a result, The norm for the numerical solutions of Richards’ equation
cell-centered finite differences and low-order finite volumes/ is low-order time discretizations, which are globally first-order.
integrated finite differences, and piecewise linear, continuous Notable exceptions include the approaches in (Tocci et al.,
Galerkin finite element schemes remain the dominant schemes 1997; Farthing et al., 2003b) based on higher order Backward
used in practice (Selker and John, 2004; Kollet and Maxwell, 2006; Difference Formulas (BDFs) and the second-order Taylor-
Yeh et al., 2011; Diersch, 2013). We note that post-processing is Gladwell scheme introduced in (Kavetski et al., 2001a, 2004).
increasingly being used in conjunction with continuous Galerkin The vast majority of temporal approximations are also implicit
methods to provide velocity fields that are locally conservative in head-based models that allow fully saturated conditions
over elemental control volumes (Larson and Niklasson, 2004; Sun to develop in the domain (Celia et al., 1990). Fully implicit
and Wheeler, 2006; Kees et al., 2008; Scudeler et al., 2016).
Local features like steep infiltration fronts make Richards’
equation a natural candidate for adaptive spatial discretizations

Fig. 3. Three widely used spatial discretizations shown for example


Fig. 2. Example piecewise linear weighting function. in 1-dimension.

www.soils.org/publications/sssaj 1261
discretizations incorporate the current solution at a time level solving a fixed number of linear systems per time step (Paniconi et
into their approximations and so require solution of at least one al., 1991; Kavetski et al., 2002). Linearly implicit approximations
nonlinear problem at each time step. can be found in either multistep or multistage (e.g., Rosenbrock
To make the following discussion more precise, we assume a schemes) contexts (Farthing et al., 2006). However they are still
method of lines (MOL) approach to eliminate partial derivative much less common than fully implicit techniques.
terms after applying one of the spatial discretizations from Adaptive time discretizations for Richards’ equation are
Section 3. The subsequent discussion would be analogous if much more common than adaptive spatial discretizations. A
we followed a Rothe paradigm and introduced a temporal classical approach is to increase or decrease the time step by a fixed
discretization before the spatial discretization (Lang, 2013). fraction based on the number of iterations taken by the nonlinear
Applying the MOL leads to following semi-discrete problem solver (Paniconi and Putti, 1994). The methods introduced in
when stepping forward from one time level, tn, to the next, tn+1 (Kavetski et al., 2001a, 2001b) and (Tocci et al., 1997) adapt the
time step size and/or order based on estimates of temporal error.
, y , y ′) 0, for t∈(t n , t n+1 ) [9]
F(t= These approaches combine low and high-order approximations
to establish estimates of the local error associated with a solution
Here, yʹÎ Rn is an n-dimensional vector of discrete unknowns using a given time step and approximation order. They accept
or degrees of freedom resulting from the spatial approximation, or reject a solution based on user-selected tolerances and pick a
y¢ is the time derivative for y, and the time step is Δt(n+1). Initial new order and/or time step size using heuristics often motivated
conditions for Eq. [9] are taken from time level n, and we assume by simple control theory (Söderlind, 2006). These approaches
that they are consistent can be orders of magnitude more efficient than fixed time step
or heuristic alternatives, depending on the level of temporal
F(t n , y n , y ′, n ) = 0 [10] accuracy requested (Brenan et al., 1996; Kees and Miller, 2002;
Farthing et al., 2003b). On the other hand, heuristic time
Equation [9] is known as a system of Differential Algebraic stepping based on nonlinear solver performance may outperform
Equations (DAEs) and can be thought of as a generalization of adaptation driven by local truncation error estimates for lower
an Ordinary Differential Equation (ODE) system (Brenan et al., accuracy regimes (D’Haese et al., 2007).
1996). As an example, BDFs are multistep methods that can be It is common to adjust time steps based on nonlinear solver
thought of as introducing an approximation for y¢, convergence as well as temporal truncation error estimates in
mature ODE and DAE solvers (Brenan et al., 1996; Hairer et al.,
y ′, n+1 ≈a n+1 y n+1 + b n [11] 1996). However, traditional approaches are not always well-suited
to highly nonlinear problems, especially when low temporal
and so lead to a discrete nonlinear problem accuracy is acceptable (Farthing et al., 2003b; Söderlind and
Wang, 2006). Efforts to combine nonlinear solver performance
F ( t n+1 , y=
n+1
, y ′, n+1 ) G=
n+1
( y n+1 ) 0 [12] and formal truncation error include Gustafsson and Söderlind
(1997), however, such approaches have not been evaluated
that must be solved to obtain y(n+1). Note that Backward Euler is thoroughly for Richards’ equation and are not common practice.
just the first order BDF
1 1 Linear and Nonlinear Solvers
y ' , n+1 ≈ n+1
y n+1 − n+1 y n [13] The desire to adjust a time step based on the performance of
Dt Dt
the nonlinear solver arises from the fundamentally local nature of
Similarly, an implicit k stage Runge-Kutta approximation for Eq. Newton’s method and its variants that serve as the linearization
[9] requires solution of the nonlinear problems (Brenan et al., techniques for the vast majority of Richards’ equation solvers.
1996; Hairer et al., 1996) That is, returning to the notation for the discrete MOL system
at time level tn+1
( k
)
F t in+1 , y n +D t n+1 ∑ j=1 aij Y ' j , Y ' i =
0, for i =
1,..., k [14]
G n+1 ( y n+1 ) = 0 [16]
for the intermediate stage derivatives Y¢i to obtain
Newton’s method attempts to solve Eq. [16] through a series of
y n+=1 y n + ∑ j=1 b j Y ' j [15]
k

linear problems
Here the implicitness results from one or more coefficient aij >
0, j ≥ i. −G n+1 ( y ni +1 ) , y ni++11 =
J ni +1D y ni +1 = y ni +1 +D y ni +1 , i =
1,... [17]
The nonlinear system of equations in Eq. [12] or [14] must,
∂G
in general, be solved iteratively until some error tolerance is met. until some convergence tolerance is met. Here J ni +1 =
∂ y
( y ni +1 )
On the other hand, there has been interest over the years in linearly is the Jacobian evaluated at y ni+1 Common convergence criteria
implicit or non-iterative temporal approximations, which require include specifying tolerances on the norm of the residual,

1262 Soil Science Society of America Journal


Gn+1(yn+1), or solution increment, Δyn+1. Classical convergence system. For large problems, the linear systems that must be solved
results show that Newton’s method for Eq. [16] will converge at each nonlinear iteration must be solved iteratively as well. The
quadratically assuming that Gn+1 is sufficiently smooth and the most frequently used iterative linear solvers are Krylov based.
initial guess, y ni+1 is sufficiently close to the true solution (Kelley, Briefly, Krylov subspace methods for the linear system Ax = b
1995). As a result, a smaller time step, Δtn+1, often translates into search for a solution at iteration j in the Krylov space,
a better initial guess for Gn+1 and improved convergence. An
added benefit of smaller time steps is that the diagonal terms of K j = span ( r0 , Ar0 , A 2 r0 , ..., A j−1r0 ) [18]
J are frequently proportional to 1⁄Δtn+1, and the linear system in
Eq. [17] becomes easier to solve as these diagonal terms become where r0 = b – Ax0. Conjugate Gradients are the canonical Krylov
more dominant (Söderlind, 2006). method and can be related to a minimization problem over Κj
Picard iteration and modified Picard iteration can be recast for a symmetric, positive definite A. For non-symmetric matrices,
as Newton’s method with an approximate Jacobian (Paniconi which is typically the case with J unless Picard iteration is used,
and Putti, 1994) and so inherit this local nature. Like other methods like Generalized Minimal RESidual (GMRES) or Bi-
modified and inexact Newton’s methods (Kelley, 1995), they Conjugate Gradient STABilized (BiCGSTAB) are required (Saad
trade reduced convergence rates in hope of gains in simplicity, and van der Vorst, 2000). The time spent in solution of this system
efficiency, and possibly robustness since the need for derivatives is usually the dominant expense in the linearization process for
of nonlinear soil constitutive relations is either eliminated large problems (Tocci et al., 1998; Jones and Woodward, 2001)
(classical Picard) or reduced (modified Picard) (Lehmann and A similar need to manage the interplay between the
Ackerer, 1998). Another approach to reduce the impact of temporal approximation, initial guess, and nonlinear problem,
highly nonlinear constitutive relations on Newton convergence G(n+1) = 0, plays out in the resulting linearized problem through
is to switch the primary variable between water content in drier the conditioning of J. That is, direct (e.g., LU-based) solvers
regions and capillary head elsewhere (Forsyth et al., 1995). are essentially global and seldom fail unless the linear problem
An interesting alternative to Newton or Picard methods is is extremely ill-conditioned. Sparse direct solvers saw much
the relaxation scheme used in Pop et al. (2004) as an approach improvement during the previous decade (Hénon et al., 2002; Li,
for linearizing mixed finite element approximations for Richards’ 2005) and are practical for many two-dimensional problems and
equation. It introduces a global, regularizing term into the moderate three-dimensional problems (Farthing et al., 2003a).
linearized problem for Gn+1 and exploits the monotonicity of soil- However, they are still unable to handle large, coupled three-
moisture constitutive relationships to secure robust convergence dimensional problems where iterative, Krylov-based techniques
behavior with potentially less restrictive time step constraints are required (Knoll and Keyes, 2004). The performance of Krylov
and better conditioned linear systems than Newton or modified methods in turn depends largely on effective preconditioning
Picard approaches (Slodicka, 2002; List and Radu, 2016). techniques. There have been efforts to develop or apply robust,
The issues associated with the local nature of Newton’s scalable preconditioners for Richards’ equation (Jenkins et al., 2001;
method and its variants are not unique to Richards’ equation and Jones and Woodward, 2001), and high-quality software packages
the need for additional modifications (so-called globalization like PETSc provide access to a number of modern preconditioners
techniques) like line-search and trust-region methods (Kelley, as well as linear and nonlinear solvers (Balay et al., 1997). However,
1995; Jones and Woodward, 2001; Knoll and Keyes, 2004; uniformly robust preconditioning techniques and nonlinear solver
Wang and Tchelepi, 2013) has been recognized in the broader performance for Richards’ equation for remains an open research
community. There has also been renewed interest in acceleration topic (Juncu et al., 2011; Lipnikov et al., 2016).
techniques like Anderson acceleration (Walker and Ni, 2011), For example, rather than attempt to extend or modify
which can be seen as an approach to improve the quality of existing preconditioners like geometric and algebraic multigrid
Picard (fixed-point) iterates by exploiting the nonlinear residual to the linearized systems arising from standard discretizations
history using Krylov subspace techniques (Lott et al., 2012). of Richards’ equation, the work of Berninger et al. (2011) takes
Finally, a globalization approach that has been considered by an alternate view and modifies the underlying discretization to
several researchers is to marry Newton’s method with a slower realize the performance of multigrid for classical linear self-adjoint
converging technique like Picard iteration (Lehmann and problems. That is, introducing a Kirchoff (or matric flux potential)
Ackerer, 1998; Bergamaschi et al., 1999) or the scheme from transform for homogeneous media (Williams et al., 2000)
Pop et al. (2004) to improve robustness for poor initial guesses
y
while recapturing quadratic convergence as intermediate iterates u ( y ) = ∫ K ( p ) dp [19]
−∞
approach the true solution (List and Radu, 2016).
As discussed above, the choice of time step naturally impacts has the effect of linearizing the gradient term in Darcy’s law
nonlinear solver performance (Gustafsson and Söderlind, through the chain rule Ñu = K(y)Ñy. Together with temporally
1997), and Richards’ equation solvers that use an implicit time explicit treatment of gravitational terms, this produces a system
discretization must negotiate a balance between the size of time that is equivalent to a convex minimization problem and for
step that can be taken and the difficulty of the resulting nonlinear which a class of nonlinear multigrid techniques are well-behaved

www.soils.org/publications/sssaj 1263
(Kornhuber, 2002). While this suffices for homogeneous the proper location of high-resolution space should also improve
domains, additional conditions based on conserving mass robustness and accuracy in many cases.
(via flux matching) and maintaining continuity of pressure at Similarly, many of the issues with robustness of Richards’
material interfaces are required for heterogeneous domains. The equation solvers may be addressed with better understanding
nonlinearity of these transmission conditions also requires a of the dynamic interplay between temporal approximation and
more complex, nonlinear substructuring domain-decomposition linearization techniques. As discussed above, researchers have
approach (Berninger et al., 2014). tried to manage these dynamics through the time step (Tocci
Finally, for large problems with iterative solvers there is et al., 1997; Kavetski et al., 2002; D’Haese et al., 2007) and/or
the need to manage not just the impact of the time step on the linearization technique (Paniconi and Putti, 1994; Lehmann and
difficulty of the nonlinear system, but also its impact on the Ackerer, 1998; Lipnikov et al., 2016). Alternative linearization
resulting linear problems, which are often better conditioned for approaches (e.g., List and Radu, 2016) and better algorithms for
smaller time steps. Moreover, clear efficiency gains are possible managing the complex dynamics of adaptive solvers taken, for
by coordinating the linear and nonlinear solver tolerances example, from control-theory (Bhaya and Kaszkurewicz, 2006)
dynamically to avoid over-solving intermediate linear systems could directly improve existing production codes and also lower
(Tocci et al., 1998; Eisenstat and Walker, 1996). the barrier to adoption of more accurate spatial approximations.
We believe the notions of computational stability and
Discussion and Alternatives proportionality are useful concepts to guide the development
Computational inefficiency and lack of robustness in of modern Richards’ equation solvers. Simulators for Richards’
Richards’ equation solvers used in practice most often manifest equation should have clearly defined input parameters for
as poor linear and nonlinear solver convergence and/or poor controlling their accuracy, and changes in these parameters should
performance of the time integrator. Our view is that a significant reliably increase (or decrease) accuracy and computational expense
portion of this poor convergence and instability arises from at a predictable rate. In other words, tightening an error tolerance
under resolution in space of sharp fronts or other phenomena slightly should lead to a commensurate decrease in error without
where highly nonlinear soil properties change drastically over excessive increases in run time (Söderlind and Wang, 2006).
neighboring elements or cells. However, the spatial component The idea of computational stability follows the line of
of this behavior is not as readily apparent because estimates or introducing mathematical concepts like continuous data
indicators of spatial accuracy are more difficult to obtain than dependence and well-posedness to computational software. The
the convergence estimates and temporal error indicators that are need for more rigorous computational science with standards
available for classical iterative linear and nonlinear techniques like reproducibility and repeatability of numerical experiments
and many time discretizations. Indeed, as discussed above, the has been recognized across a number of fields (Collberg and
vast majority of spatial discretizations used in practice are based Proebsting, 2016). It can be seen, for example, in efforts to
on classical low-order finite element discretizations and cell- develop clear understanding of benchmarking, verification,
centered finite differences and are combined with evaluation and validation for computational codes in fields like nuclear
techniques for soil constitutive relations like mass-lumping and engineering and aerospace engineering (Babuska and Oden,
upwinding that are designed to favor robustness over accuracy. 2004; Oberkampf and Trucano, 2007). A key component of
One way to increase computational efficiency at the such efforts is a suite of test problems identified and accepted
expense of solution validity is to use resolutions considerably by the community that allows fair evaluation and comparison
coarser than the REV scale dictates. This requires the use of so- of numerical techniques and simulation codes. While several
called “effective” parameter values that diverge from physical test problems recur throughout the literature and isolated
reality as the computational scale increases, and results in large comparisons can be found, the community has lacked a central
errors in the simulation’s representation of the physical system repository of benchmark problems or recurring efforts to compare
even though it has converged. We hope instead that there will results across research groups as can be found in other fields
be continued development and adoption of new spatial and (e.g., American Institute of Aeronautics and Astronautics, 1998;
temporal discretization schemes. Following trends of maturing Larson et al., 2014; American Society of Mechanical Engineers,
computational fields, we expect to see improved, adaptive 2017; International Association for the Engineering Modelling,
algorithms driven by solution dynamics and a posteriori error Analysis and Simulation Community, http://www.nafems.
estimates (Mostaghimi et al., 2015; Baron et al., 2017). Moreover, org [verified 14 Sept. 2017]). The recent integrated hydrologic
approaches are needed that consider entire solution error and model comparison project (Maxwell et al., 2014; Kollet et al.,
dynamics–for example, joint space-time discretization and error 2017) and the International Soil Modeling Consortium (http://
control (Solin and Kuraz, 2011) rather than just a MOL approach soil-modeling.org; verified 10 Feb. 2017) represent welcome
with temporal adaptivity but static spatial approximation steps in this direction. This is also an area where government
(Farthing et al., 2003b). While the implementation of full agencies (e.g., NOAA, the US National Weather Service, Office
spatially and temporally adaptive techniques is a clear hurdle, of Water Prediction) could make a major contribution.

1264 Soil Science Society of America Journal


Even as we work to improve the numerics, there continues the vadose zone into N regions of thickness Δθ called “bins”
to be reason to evaluate Richards’ equation as a formulation (Talbot and Ogden, 2008).
for unsaturated flow. Problems where gas-phase dynamics That resulting ODE for infiltrating water is (Ogden et
are important usually demand a full two-phase formulation al., 2015a):
(Helmig, 2011; Smits et al., 2013), while more recent work on
 dz  K ( q d ) − K ( q i )  y ( q d ) + hp 
thermodynamically constrained averaging theory has =focused    + 1  [21]
dt
 j q − q z 
on developing a consistent upscaling framework for deriving
d i  j 
extensions to Richards’ equation (Bronson, 2014; Gray and where j is the bin index, θd is the moisture content of the bin
Miller, 2014). These approaches are primarily aimed at extending containing the highest moisture content, θi is the initial moisture
the operational range of validity for the mathematical system content, hp is the ponded depth if existent, and zj is the depth
as a representation for unsaturated flow (Or et al., 2015). In to the wetting front in the jth finite moisture content bin. Note
some cases, new formulations or closure schemes may introduce that the ODE solution of Eq. [21] requires evaluation of no
additional challenges for numerical schemes due to new terms spatial derivatives, eliminating what is a source of difficulty in
with additional, nonlinear coupling and complexity. On the the traditional solution of Richards’ equation. Equation [21]
other hand, it is also possible that some numerical difficulties are can be solved using an explicit forward Euler scheme or other
unintended side effects of inadequate formulations rather than traditional ODE solutions such as Runge-Kutta. This ODE
being inherent to the unsaturated flow problem. is guaranteed to converge and to conserve mass using a finite
In a different direction, recent work to derive alternative volume solution. While the SMVE solution method is at present
formulations that trade some of the generality of the traditional limited to one-dimensional infiltration in homogeneous soil
Richards’ equation solution for gains in speed and reliability has layers, the reliable, accurate and efficient ODE solution are major
shown success. It is possible, for example, to convert Richards’ advantages. Furthermore, Eq. [21] will solve sharp wetting fronts
equation into a formulation for the evolution of soil moisture with no difficulty and is not degenerate, making the method well
front location with θ as the independent variable under suitable suited for use in Earth System and large-scale models of land-
assumptions (Philip, 1957; Swartzendruber, 1969). Ogden et al. atmosphere interaction.
(2017) converted the one-dimensional Richards’ equation into a
form that they call the Soil Moisture Velocity Equation (SMVE) Conclusions
that describes the speed of propagation of particular moisture We provided a high-level overview of numerical methods
contents within a homogeneous soil layer: for the solution of Richards’ equation with an eye toward
summarizing previous work and identifying major trends. A
 dz  −∂ K ( q )  ∂ y ( q )  ∂ 2 y/∂ z 2 [20]
=    −1  − D ( )
q recurring theme is the need to balance speed, robustness, and
 dt  q ∂q  ∂z  ∂ y/∂ z accuracy. Historically, the dominant approaches have been low
Equation [20] is equivalent to the one-dimensional Richards order in space and time and used fixed spatial grids or meshes. This
equation. This equation separates the flux due to gravity and the has set up an unfortunate tension in many cases between accuracy
integrated wetting front capillary drive from the flux caused and robustness– or at least ease of convergence. Loosening
by the shape of the wetting front. The first term on the right- tolerances and coarsening resolution are expedient, but unreliable,
hand side Eq. [20], which is called the “advection-like” term, was ways to improve speed or achieve convergence that have a negative
derived by Ogden et al. (2015a). The second term on the right- effect on solution accuracy and realism. The pursuit of improved
hand side of Eq. [20], called the “diffusion-like” term, is equal to numerical methods has largely been about reducing this tension
a diffusive flux due to the second spatial derivative of capillary and increasing computational efficiency rigorously.
head divided by the slope of the capillary head along the wetting We hope to see continued development and research into
front. This diffusion-like term will be zero when the gradient of the numerical solution of Richards’ equation. Indeed, we believe
the capillarity is a constant along lines of constant θ, or in the it is essential if application of Richards’ equation to large-scale
case of a sharp wetting front when ¶y/¶z = ¥. Both of these problems at valid discretizations is to become commonplace.
cases represent situations that can cause convergence difficulties Successful Richards’ equation solvers in the future will likely
in Richards’ equation solvers. be adaptive in space and time, while efficiently managing the
Fluxes calculated using the SMVE advection-like term agree interplay between discretizations on the one hand and linear
very closely with the numerical solution of Richards’ equation and/or nonlinear algebra solvers on the other. The recent
(Ogden et al., 2015a), data from column experiments (Ogden et identification of the Soil Moisture Velocity Equation (Eq. [20])
al., 2015b) and exact analytical solutions (Ogden et al., 2017). indicates that, at least in the case of one-dimensional solutions
Results in all cases show that the diffusion-like term in Eq. of Richards’ equation, there may be novel approaches yet to be
[20] is negligible unless one is interested in the exact shape of discovered. The most likely way for this development to continue
the wetting front. Most importantly, the advection-like term of is to encourage the exchange of ideas and techniques with the
the SMVE can be converted into an ODE using the MOL and applied mathematics, computational mechanics, computer
solved using a finite moisture-content discretization that divides science and Earth Science communities. Improved software

www.soils.org/publications/sssaj 1265
saturated groundwater flow. Vadose Zone J. 8(2):352–362. doi:10.2136/
engineering and creation of archival verification and validation
vzj2008.0108
data sets for use by the community are also essential if we are to Bergamaschi, L., B. Luca, and M. Putti. 1999. Mixed finite elements and
proceed rationally and make the most of our efforts. Newton‐type linearizations for the solution of Richards equation. Int.
J. Numer. Methods Eng. 45(8):1025–1046. doi:10.1002/(SICI)1097-
0207(19990720)45:8<1025::AID-NME615>3.0.CO;2-G
Acknowledgments Berninger, H., B. Heiko, K. Ralf, and S. Oliver. 2014. A multidomain
Permission was granted by the Chief of Engineers to publish this discretization of the Richards equation in layered soil. Comput. Geosci.
information. This material is based in part on work supported by 19(1):213–232. doi:10.1007/s10596-014-9461-8
the National Science Foundation under Grant EAR-1360384 to the Berninger, H., R. Kornhuber, and O. Sander. 2011. Fast and robust numerical
University of Wyoming. Any opinions, findings, and conclusions or solution of the Richards equation in homogeneous soil. SIAM J. Numer.
recommendations expressed in this material are those of the author(s) and Anal. 49(6):2576–2597. doi:10.1137/100782887
do not necessarily reflect the views of the National Science Foundation. Bhaya, A., and E. Kaszkurewicz. 2006. Control perspectives on numerical
algorithms and matrix problems. Society for Industrial and Applied
Mathematics, Philadelphia, PA. doi:10.1137/1.9780898718669
References Brenan, K.E., S.L. Campbell, and L.R. Petzold. 1996. Numerical solution
Aavatsmark, I., T. Barkve, O. Bøe, and T. Mannseth. 1998. Discretization of initial-value problems in differential-algebraic equations.
on unstructured grids for inhomogeneous, anisotropic media. Part I: Society for Industrial and Applied Mathematics, Philadelphia, PA.
Derivation of the methods. SIAM J. Sci. Comput. 19(5):1700–1716. doi:10.1137/1.9781611971224
doi:10.1137/S1064827595293582 Broadbridge, P., and I. White. 1988. Modelling solute transport, chemical
Allen, M.B., and C. Murphy. 1985. A finite element collocation method for adsorption and cation exchange. In: Int. Hydrology and Water Resources
variably saturated flows in porous media. Numer. Methods Partial Differ. Symp. Nat. Conf. Publ. No. 92/19 (Preprints of Papers, Vol. 3:924-929).
Equ. 1:229–239. doi:10.1002/num.1690010306 2-4 October. Institution of Engineers, Perth, Australia.
Alt, W.H., and S. Luckhaus. 1983. Quasilinear elliptic-parabolic differential Bronson, S. 2014. Assessing the minimum entropy production rate principle
equations. Math. Z. 183:311–341. doi:10.1007/BF01176474 for multiphase flow using the thermodynamically constrained averaging
American Institute of Aeronautics and Astronautics. 1998. AIAA guide for the theory approach. University of North Carolina, Chapel Hill.
verification and validation of computational fluid dynamics simulations. Brunner, P., and C.T. Simmons. 2012. HydroGeoSphere: A fully integrated,
American Institute of Aeronautics & Astronautics, Reston, VA. physically based hydrological model. Ground Water 50:170–176.
American Society of Mechanical Engineers. 2017. ASME V&V, Verification and doi:10.1111/j.1745-6584.2011.00882.x
validation symposium. American Society of Mechanical Engineers, Las Campbell, J.C., J.M. Hyman, and M.J. Shashkov. 2002. Mimetic finite difference
Vegas, NV. 3–5 May 2017. https://www.asme.org/events/vandv (verified operators for second-order tensors on unstructured grids. Comput. Math.
6 Sept. 2017). Appl. 44(1-2):157–173. doi:10.1016/S0898-1221(02)00137-2
Arbogast, T., A. Todd, C.N. Dawson, P.T. Keenan, M.F. Wheeler, and Y. Ivan. Camporese, M., C. Paniconi, M. Putti and S. Orlandini. 2010. Surface-subsurface
1998. Enhanced cell-centered finite differences for elliptic equations on flow modeling with path-based runoff routing, boundary condition-based
general geometry. SIAM J. Sci. Comput. 19(2):404–425. doi:10.1137/ coupling, and assimilation of multisource observation data. Water Resour.
S1064827594264545 Res. 46. doi:10.1029/2008WR007536.
Arbogast, T., A. Todd, and C. Zhangxin. 1995. On the implementation of mixed Celia, M.A., E.T. Bouloutas, and R.L. Zarba. 1990. A general mass-conservative
methods as nonconforming methods for second-order elliptic problems. numerical solution for the unsaturated flow equation. Water Resour. Res.
Math. Comput. 64(211):943–972. 26(7):1483–1496. doi:10.1029/WR026i007p01483
Arrarás, A., L. Portero, and J.C. Jorge. 2009. A combined mixed finite element Chavent, G., G. Cohen, and J. Jaffre. 1984. Discontinuous upwinding and mixed
ADI scheme for solving Richards’ equation with mixed derivatives on finite elements for two-phase flows in reservoir simulation. Computer
irregular grids. Appl. Numer. Math. 59(3-4):454–467. doi:10.1016/j. Methods in Applied Mechanics and Engineering 47:93–118.
apnum.2008.03.002 Chavent, G., and J.E. Roberts. 1991. A unified physical presentation of mixed,
Assouline, S. 2013. Infiltration into soils: Conceptual approaches and solutions. mixed-hybrid finite elements and standard finite difference approximations
Water Resour. Res. 49:1755–1772. doi:10.1002/wrcr.20155 for the determination of velocities in waterflow problems. Adv. Water
Babuska, I., and J.T. Oden. 2004. Verification and validation in computational Resour. 14(6):329–348. doi:10.1016/0309-1708(91)90020-O
engineering and science: Basic concepts. Comput. Methods Appl. Mech. Chen, Z., G. Huan, and Y. Ma. 2006. Computational methods for multiphase
Eng. 193(36-38):4057–4066. doi:10.1016/j.cma.2004.03.002 flows in porous media. Society for Industrial and Applied Mathematics,
Balay, S., W.D. Gropp, L.C. McInnes, and B.F. Smith. 1997. Efficient management Philadelphia, PA. doi:10.1137/1.9780898718942
of parallelism in object-oriented numerical software libraries. In: E. Arge, Collberg, C., and T.A. Proebsting. 2016. Repeatability in computer systems
A.M. Bruaset, and H.P. Langtengen, editors, Modern software tools for research. Commun. ACM 59(3):62–69.
scientific computing. Springer, New York. p. 163–202. Cox, P.M., R.A. Betts, C.B. Bunton, R.L.H. Essery, P.R. Rowntree, and J. Smith.
Baron, V., Y. Coudière, and P. Sochala. 2017. Adaptive multistep time 1999. The impact of new land surface physics on the GCM simulation
discretization and linearization based on a posteriori error estimates for of climate and climate sensitivity. Clim. Dyn. 15:183–203. doi:10.1007/
the Richards equation. Applied Numerical Mathematics 112:104–125. s003820050276
doi:10.1016/j.apnum.2016.10.005 Dawson, C., S. Sun, and M.F. Wheeler. 2004. Compatible algorithms for
Barry, D.A., and G.C. Sander. 1991. Exact solutions for water infiltration with coupled flow and transport. Comput. Methods Appl. Mech. Eng. 193(23-
an arbitrary surface flux or nonlinear solute adsorption. Water Resour. Res. 26):2565–2580. doi:10.1016/j.cma.2003.12.059
27:2667–2680. doi:10.1029/91WR01445 D’Haese, C.M.F., M. Putti, C. Paniconi, and N.E.C. Verhoest. 2007. Assessment
Barry, D.A., J.-Y. Parlange, G.C. Sander, and M. Sivapalan. 1993. A class of exact of adaptive and heuristic time stepping for variably saturated flow. Int. J.
solutions for Richards’ equation. J. Hydrol. 142:29–46. doi:10.1016/0022- Numer. Methods Fluids 53(7):1173–1193. doi:10.1002/fld.1369
1694(93)90003-R Diersch, H.J. 2013. Flow in variably saturated porous media. In: H.J. Diersch,
Bartlett, R.A. 2009. Integration strategies for computational science and editor, FEFLOW: Finite Element Modeling of Flow, Mass and Heat
engineering software. In: 2009 ICSE Workshop on Software Engineering Transport in Porous and Fractured Media. Springer Verlag. Heidelberg,
for Computational Science and Engineering. Vancouver, BC, Canada. 23- Germany.
23 May 2009. doi:10.1109/SECSE.2009.5069160 Downer, C.W., and F.L. Ogden. 2004. Appropriate vertical discretization of
Bause, M., and P. Knabner. 2004. Computation of variably saturated subsurface Richards’ equation for two-dimensional watershed-scale modelling.
flow by adaptive mixed hybrid finite element methods. Adv. Water Resour. Hydrol. Processes 18:1–22. doi:10.1002/hyp.1306
27(6):565–581. doi:10.1016/j.advwatres.2004.03.005 Eisenstat, S.C., and H.F. Walker. 1996. Choosing the forcing terms in
Belfort, B., B. Benjamin, R. Fanilo, Y. Anis, and L. François. 2009. An an inexact Newton method. SIAM J. Sci. Comput. 17(1):16–32.
efficient lumped mixed hybrid finite element formulation for variably

1266 Soil Science Society of America Journal


doi:10.1137/0917003 media. Comput. Geosci. 4:141–164. doi:10.1023/A:1011559916309
Ek, M.B., K.E. Mitchell, Y. Lin, E. Rogers, P. Grunmann, V. Koren, and J.D. Jenkins, E.W., C.E. Kees, C.T. Kelley, and C.T. Miller. 2001. An aggregation-
Tarpley. 2003. Implementation of Noah land surface model advances based domain decomposition preconditioner for groundwater flow. SIAM
in the National Centers for Environmental Prediction operational J. Sci. Comput. 23(2):430–441. doi:10.1137/S1064827500372274
mesoscale Eta model. J. Geophys. Res., D, Atmospheres 108(D22). Jones, J.E., and C.S. Woodward. 2001. Newton–Krylov-multigrid solvers for
doi:10.1029/2002JD003296 large-scale, highly heterogeneous, variably saturated flow problems. Adv.
Ern, A., and J.-L. Guermond. 2004. Theory and practice of finite elements. Water Resour. 24(7):763–774. doi:10.1016/S0309-1708(00)00075-0
Springer Verlag, Cambridge. doi:10.1007/978-1-4757-4355-5 Juncu, G., J. Gheorghe, N. Aurelian, and P. Constantin. 2011. Nonlinear
Farthing, M.W., C.E. Kees, T.S. Coffey, C.T. Kelley, and C.T. Miller. 2003a. multigrid methods for numerical solution of the variably saturated flow
Efficient steady-state solution techniques for variably saturated equation in two space dimensions. Transp. Porous Media 91(1):35–47.
groundwater flow. Adv. Water Resour. 26(8):833–849. doi:10.1016/ doi:10.1007/s11242-011-9831-9
S0309-1708(03)00076-9 Kavetski, D., P. Binning, and S.W. Sloan. 2001a. Adaptive time stepping and
Farthing, M.W., C.E. Kees, and C.T. Miller. 2003b. Mixed finite element error control in a mass conservative numerical solution of the mixed form
methods and higher order temporal approximations for variably saturated of Richards equation. Adv. Water Resour. 24(6):595–605. doi:10.1016/
groundwater flow. Adv. Water Resour. 26(4):373–394. doi:10.1016/ S0309-1708(00)00076-2
S0309-1708(02)00187-2 Kavetski, D., P. Binning, and S.W. Sloan. 2001b. Adaptive backward Euler
Farthing, M.W., C.E. Kees, E.W. Jenkins, and C.T. Miller. 2006. An evaluation of time stepping with truncation error control for numerical modelling of
linearly implicit time discretization methods for approximating Richards’ unsaturated fluid flow. Int. J. Numer. Methods Eng. 53(6):1301–1322.
equation. In: Binning, P., Engesgaard, P., Dahle, H., Pinder, G., Gray, W.G., doi:10.1002/nme.329
editors, XVI International Conference on Computational Methods in Kavetski, D., P. Binning, and S.W. Sloan. 2002. Noniterative time stepping
Water Resources (CMWR-XVI). Technical University of Denmark. schemes with adaptive truncation error control for the solution
Fidkowski, K.J., and D.L. Darmofal. 2011. Review of output-based error of Richards equation. Water Resour. Res. 38(10):29-1–29-10.
estimation and mesh adaptation in computational fluid dynamics. AIAA doi:10.1029/2001WR000720
J. 49(4):673–694. doi:10.2514/1.J050073 Kavetski, D., P. Binning, and S.W. Sloan. 2004. Truncation error and stability
Finsterle, S., C. Doughty, M.B. Kowalsky, G.J. Moridis, L. Pan, T. Xu, et al. 2008. analysis of iterative and non-iterative Thomas–Gladwell methods for
Advanced vadose zone simulations using TOUGH. Vadose Zone J. 7:601– first-order non-linear differential equations. Int. J. Numer. Methods Eng.
609. doi:10.2136/vzj2007.0059 60(12):2031–2043. doi:10.1002/nme.1035
Forsyth, P.A., and M.C. Kropinski. 1997. Monotonicity considerations for Kees, C.E., M.W. Farthing, and C.N. Dawson. 2008. Locally conservative,
saturated–unsaturated subsurface flow. SIAM J. Sci. Comput. 18(5):1328– stabilized finite element methods for variably saturated flow. Comput.
1354. doi:10.1137/S1064827594265824 Methods Appl. Mech. Eng. 197(51-52):4610–4625. doi:10.1016/j.
Forsyth, P.A., Y.S. Wu, and K. Pruess. 1995. Robust numerical methods for cma.2008.06.005
saturated-unsaturated flow with dry initial conditions in heterogeneous Kees, C.E., and C.T. Miller. 2002. Higher order time integration methods for
media. Adv. Water Resour. 18(1):25–38. doi:10.1016/0309- two-phase flow. Adv. Water Resour. 25(2):159–177. doi:10.1016/S0309-
1708(95)00020-J 1708(01)00054-9
Gerritsen, M.G. and L.J. Durlofsky. 2005. Modeling fluid flow in oil reservoirs. Kelley, C.T. 1995. Iterative methods for linear and nonlinear equations. Society
Annual Review of Fluid Mechanics 37:211–238. for Industrial and Applied Mathematics, Philadelphia, PA.
Gilding, B.H. 1991. Qualitative mathematical analysis of the Richards equation. Kirkland, M.R., R.G. Hills, and P.J. Wierenga. 1992. Algorithms for solving
Transp. Porous Media 6(5-6):651–666. doi:10.1007/BF00137854 Richards’ equation for variably saturated soils. Water Resour. Res.
Gray, W.G., and S. Hassanizadeh. 1991. Paradoxes and realities in unsaturated flow 28(8):2049–2058. doi:10.1029/92WR00802
theory. Water Resour. Res. 27(8):1847–1854. doi:10.1029/91WR01259 Klausen, R.A., and T.F. Russell. 2004. Relationships among some locally
Gray, W.G., and C.T. Miller. 2014. Introduction to the thermodynamically conservative discretization methods which handle discontinuous
constrained averaging theory for porous medium systems. Springer, coefficients. Comput. Geosci. 8(4):341–377. doi:10.1007/s10596-005-
Cambridge. doi:10.1007/978-3-319-04010-3 1815-9
Gustafsson, K., and G. Söderlind. 1997. Control strategies for the iterative Klausen, R.A., F.A. Radu, and G.T. Eigestad. 2008. Convergence of MPFA on
solution of nonlinear equations in ODE solvers. SIAM J. Sci. Comput. triangulations and for Richards’ equation. Int. J. Numer. Methods Fluids
18(1):23–40. doi:10.1137/S1064827595287109 58(12):1327–1351. doi:10.1002/fld.1787
Hairer, E., S.P. Nørsett, and G. Wanner. 1996. Solving ordinary differential Knoll, D.A., and D.E. Keyes. 2004. Jacobian-free Newton–Krylov methods: A
equations I: Nonstiff problems. Second ed. Springer Verlag, Cambridge. survey of approaches and applications. J. Comput. Phys. 193(2):357–397.
doi:10.1007/978-3-642-05221-7 doi:10.1016/j.jcp.2003.08.010
Healy, R.W. 2008. Simulating water, solute, and heat transport in the Kollet, S.J., and R.M. Maxwell. 2006. Integrated surface–groundwater
subsurface with the VS2DI software package. Vadose Zone J. 7:632–639. flow modeling: A free-surface overland flow boundary condition in a
doi:10.2136/vzj2007.0075 parallel groundwater flow model. Adv. Water Resour. 29(7):945–958.
Helmig, R. 2011. Multiphase flow and transport processes in the subsurface: A doi:10.1016/j.advwatres.2005.08.006
contribution to the modeling of hydrosystems. Springer, Cambridge. Kollet, S., M. Sulis, R.M. Maxwell, C. Paniconi, M. Putti, G. Bertoldi, E.T. Coon,
Hénon, P., P. Ramet, and J. Roman. 2002. PaStiX: A high-performance parallel E. Cordano, S. Endrizzi, E. Kikinzon, E. Mouche, C. Mügler, Y.-J. Park,
direct solver for sparse symmetric positive definite systems. Parallel J.C. Refsgaard, S. Stisen, and E. Sudicky. 2017. The integrated hydrologic
Comput. 28(2):301–321. doi:10.1016/S0167-8191(01)00141-7 model intercomparison project, IH-MIP2: A second set of benchmark
Heroux, M.A., and J.M. Willenbring. 2009. Barely sufficient software results to diagnose integrated hydrology and feedbacks. Water Resour. Res.
engineering: 10 practices to improve your CSE software. In: 2009 ICSE doi:10.1002/2016WR019191
Workshop on Software Engineering for Computational Science and Kornhuber, R. 2002. On constrained Newton linearization and multigrid for
Engineering. doi:10.1109/SECSE.2009.5069157 variational inequalities. Numer. Math. 91(4):699–721. doi:10.1007/
Hoteit, H., and A. Firoozabadi. 2005. Multicomponent fluid flow by s002110100341
discontinuous Galerkin and mixed methods in unfractured and fractured Kumar, M., C.J. Duffy, and K.M. Salvage. 2009. A second-order accurate, finite
media. Water Resour. Res. 41(11):W11412. doi:10.1029/2005WR004339 volume–based, integrated hydrologic modeling (FIHM) framework
Howington, S.E., R.C. Berger, J.P. Hallberg, J.F. Peters, A.K. Stagg, E.W. Jenkins, for simulation of surface and subsurface flow. Vadose Zone J. 8(4):873.
and C.T. Kelley. 1999. A model to simulate the interaction between doi:10.2136/vzj2009.0014
groundwater and surface water. ADA451802. US Army Engineering Lang, J. 2013. Adaptive multilevel solution of nonlinear parabolic PDE systems:
Research and Development Center, Vicksburg, MS. p. 1-12. Theory, algorithm, and applications. Springer Science & Business Media,
Huber, R., and R. Helmig. 2000. Node-centered finite volume discretizations Berlin, Germany.
for the numerical simulation of multiphase flow in heterogeneous porous Larson, L., F. Stern, and M. Visonneau, editors. 2014. Numerical ship

www.soils.org/publications/sssaj 1267
hydrodynamics: An assessment of the Gothenburg 2010 Workshop. Ogden, F.L., W. Lai, R.C. Steinke, and J. Zhu. 2015b. Validation of finite water-
Springer Science & Business Media, Dordrecht, Germany. content vadose zone dynamics method using column experiments with
Larson, M.G., and A.J. Niklasson. 2004. A conservative flux for the continuous a moving water table and applied surface flux. Water Resour. Res. 51.
Galerkin method based on discontinuous enrichment. Calcolo 41(2):65– doi:10.1002/2014WR016454
76. doi:10.1007/s10092-004-0084-7 Ogden, F.L., M.B. Allen, W. Lai, J. Zhu, M. Seo, C.C. Douglas, and C.A. Talbot.
Lehmann, F., and P.H. Ackerer. 1998. Comparison of iterative methods for 2017. The soil moisture velocity equation. J. Adv. Model. Earth Syst.
improved solutions of the fluid flow equation in partially saturated porous doi:10.1002/2017MS000931
media. Transport in Porous Media 31:275–292. Or, D., P. Lehmann, and S. Assouline. 2015. Natural length scales define the
Li, H., M.W. Farthing, C.N. Dawson, and C.T. Miller. 2007a. Local range of applicability of the Richards equation for capillary flows. Water
discontinuous Galerkin approximations to Richards’ equation. Adv. Water Resour. Res. 51:7130–7144. doi:10.1002/2015WR017034
Resour. 30(3):555–575. doi:10.1016/j.advwatres.2006.04.011 Orgogozo, L., N. Renon, C. Soulaine, F. Henon, S.K. Tomer, D. Labat, et al.
Li, H., M.W. Farthing, and C.T. Miller. 2007b. Adaptive local discontinuous 2014. An open source massively parallel solver for Richards equation:
Galerkin approximation to Richards’ equation. Adv. Water Resour. Mechanistic modelling of water fluxes at the watershed scale. Comput.
30(9):1883–1901. doi:10.1016/j.advwatres.2007.02.007 Phys. Commun. 185:3358–3371. doi:10.1016/j.cpc.2014.08.004
Li, X.S. 2005. An overview of SuperLU. ACM Trans. Math. Softw. 31(3):302– Pain, C.C., A.P. Umpleby, C.R.E. de Oliveira, and A.J.H. Goddard. 2001.
325. doi:10.1145/1089014.1089017 Tetrahedral mesh optimisation and adaptivity for steady-state and
Liang, X., D.P. Lettenmaier, E.F. Wood, and S.J. Burges. 1994. A simple transient finite element calculations. Comput. Methods Appl. Mech. Eng.
hydrologically based model of land surface water and energy fluxes for 190:3771–3796. doi:10.1016/S0045-7825(00)00294-2
general circulation models. J. Geophys. Res. 99(D7):14,415–14,428. Paniconi, C., P. Claudio, A.A. Aldama, and E.F. Wood. 1991. Numerical
doi:10.1029/94JD00483 evaluation of iterative and noniterative methods for the solution of the
Lipnikov, K., L. Konstantin, M. David, and S. Daniil. 2016. New preconditioning nonlinear Richards equation. Water Resour. Res. 27(6):1147–1163.
strategy for Jacobian-free solvers for variably saturated flows with doi:10.1029/91WR00334
Richards’ equation. Adv. Water Resour. 94:11–22. doi:10.1016/j. Paniconi, C., and M. Putti. 1994. A comparison of Picard and Newton
advwatres.2016.04.016 iteration in the numerical solution of multidimensional variably
List, F., and F.A. Radu. 2016. A study on iterative methods for solving Richards’ saturated flow problems. Water Resour. Res. 30(12):3357–3374.
equation. Comput. Geosci. 20:341–353. doi:10.1007/s10596-016-9566-3 doi:10.1029/94WR02046
Lott, P.A., H.F. Walker, C.S. Woodward, and U.M. Yang. 2012. An accelerated Paniconi, C., and M. Putti. 2015. Physically based modeling in catchment
Picard method for nonlinear systems related to variably saturated flow. hydrology at 50: Survey and outlook. Water Resour. Res. 51:7090–7129.
Adv. Water Resour. 38:92–101. doi:10.1016/j.advwatres.2011.12.013 doi:10.1002/2015WR017780
Manzini, G., and S. Ferraris. 2004. Mass-conservative finite volume methods Pettway, J.S., J.H. Schmidt, and A.K. Stagg. 2010. Adaptive meshing in a mixed
on 2-D unstructured grids for the Richards’ equation. Adv. Water Resour. regime hydrologic simulation model. Comput. Geosci. 14(4):665–674.
27(12):1199–1215. doi:10.1016/j.advwatres.2004.08.008 doi:10.1007/s10596-010-9179-1
Maxwell, R.M., M. Putti, S. Meyerhoff, J.-O. Delfs, I.M. Ferguson, V. Ivanov, Philip, J.R. 1957. Numerical solution of equations of the diffusion type
J. Kim, O. Kolditz, S.J. Kollet, M. Kumar, S. Lopez, J. Niu, C. Paniconi, with diffusivity concentration-dependent. II. Aust. J. Phys. 10:29–42.
Y.-J. Park, M.S. Phanikumar, C. Shen, E.A. Sudicky, and M. Sulis. 2014. doi:10.1071/PH570029
Surface-subsurface model intercomparison: A first set of benchmark Pop, I.S., F. Radu, and P. Knabner. 2004. Mixed finite elements for the Richards’
results to diagnose integrated hydrology and feedbacks. Water Resour. Res. equation: Linearization procedure. J. Comput. Appl. Math. 168(1-
50(2):1531–1549. doi:10.1002/2013WR013725 2):365–373. doi:10.1016/j.cam.2003.04.008
McConnell, S. 2004. Code Complete: A practical handbook of software Rannacher, R. 2001. Adaptive Galerkin finite element methods for partial
construction. Microsoft Press, Redmond, WA. differential equations. J. Comput. Appl. Math. 128(1-2):205–233.
Miller, C.T., C. Abhishek, and M.W. Farthing. 2006. A spatially and temporally doi:10.1016/S0377-0427(00)00513-6
adaptive solution of Richards’ equation. Adv. Water Resour. 29(4):525– Rathfelder, K., and L.M. Abriola. 1994. Mass conservative numerical solutions
545. doi:10.1016/j.advwatres.2005.06.008 of the head-based Richards equation. Water Resour. Res. 30:2579–2586.
Miller, C.T., C.N. Dawson, M.W. Farthing, T.Y. Hou, J.F. Huang, C.E. Kees, Richards, L.A. 1931. Capillary conduction of liquids through porous medium.
et al. 2013. Numerical simulation of water resources problems: Models, Physics 1:318–333.
methods, and trends. Adv. Water Resour. 51:405–437. doi:10.1016/j. Richardson, L.F. 1922. Weather prediction by numerical process. University
advwatres.2012.05.008 Press, Cambridge. p. 262.
Miller, C.T., C. George, P.T. Imhoff, J.F. McBride, J.A. Pedit, and J.A. Trangenstein. Rogers, C., M.P. Stallybrass, and D.L. Clements. 1983. On two phase filtration
1998. Multiphase flow and transport modeling in heterogeneous porous under gravity and with boundary infiltration: Application of a Backlund
media: Challenges and approaches. Adv. Water Resour. 21(2):77–120. transformation. Nonlin. Anal. Theory Meth. Appl. 7(7):785–799.
doi:10.1016/S0309-1708(96)00036-X doi:10.1016/0362-546X(83)90034-2
Mostaghimi, P., J.R. Percival, D. Pavlidis, R.J. Ferrier, J.L.M.A. Gomes, G.J. Ross, P.J., and J.-Y. Parlange. 1994. Comparing exact and numerical solutions of
Gorman, et al. 2015. Anisotropic Mesh Adaptivity and Control Volume Richards’ equation for one-dimensional infiltration and drainage. Soil Sci.
Finite Element Methods for Numerical Simulation of Multiphase Flow 157(6):341–344. doi:10.1097/00010694-199406000-00002
in Porous Media. Math. Geosci. 47:417–440. doi:10.1007/s11004-014- Saad, Y., and H.A. van der Vorst. 2000. Iterative solution of linear systems in
9579-1 the 20th century. J. Comput. Appl. Math. 123:1–33. doi:10.1016/S0377-
Niswonger, R.G., and D.E. Prudic. 2009. Comment on “Evaluating interactions 0427(00)00412-X
between groundwater and vadose zone using the HYDRUS-based flow Sander, G.C., J.-Y. Parlange, V. Kuhnel, W.L. Hogarth, D. Lockington, and J.P.J.
package for MODFLOW”. Vadose Zone J. 8:818–819. doi:10.2136/ O’Kane. 1988. Exact nonlinear solution for constant flux infiltration. J.
vzj2008.0155 Hydrol. 97:341–346. doi:10.1016/0022-1694(88)90123-0
Niu, G.-Y., Z.L. Yang, R.E. Dickinson, L.E. Gulden, and H. Su. 2007. Schwab, C. 1998. P- and Hp- finite element methods: Theory and applications in
Development of a simple groundwater model for use in climate models solid and fluid mechanics. Univ. Press, Oxford.
and evaluation with Gravity Recovery and Climate Experiment data. J. Scudeler, C., M. Putti, and C. Paniconi. 2016. Mass-conservative reconstruction
Geophys. Res. 112:D07103. doi:10.1029/2006JD007522 of Galerkin velocity fields for transport simulations. Adv. Water Resour.
Oberkampf, W.L., and T.G. Trucano. 2007. Verification and validation 94:470–485. doi:10.1016/j.advwatres.2016.06.011
benchmarks. Sandia National Laboratory, Albuquerque, NM. Selker, J., and S. John. 2004. Modelling variably saturated flow with HYDRUS-
doi:10.2172/901974 2D. Vadose Zone J. 3(2):725. doi:10.2136/vzj2004.0725
Ogden, F.L., W. Lai, R.C. Steinke, J. Zhu, C.A. Talbot, and J.L. Wilson. 2015a. Shewchuk, J.R. 2002. What is a good linear finite element?- Interpolation,
A new general 1-D vadose zone flow solution method. Water Resour. Res. Conditioning, Anisotropy, and Quality Measures. Proceedings of the 11th
51:4282–4300. doi:10.1002/2015WR017126 International Meshing Roundtable. Sandia National Laboratories and

1268 Soil Science Society of America Journal


Cornell University, Ithaca, New York. van Dam, J.C., and R.A. Feddes. 2000. Numerical simulation of infiltration,
Simunek, J., M.Th. van Genuchten, and M. Sejna. 2005. HYDRUS-1D software evaporation, and shallow groundwater levels with the Richards’ equation.
package for simulating the movement of water, heat, and multiple solutes J. Hydrol. 233(1-4):72–85. doi:10.1016/S0022-1694(00)00227-4
in variably saturated media, version 3.0, HYDRUS software series Van Duyn, C.J., and L.A. Peletier. 1982. Nonstationary filtration in partially
1. Department of Environmental Sciences, University of California saturated porous media. Arch. Ration. Mech. Anal. 78(2):173–198.
Riverside, Riverside, CA. doi:10.1007/BF00250838
Simunek, J., M.T. van Genuchten, and M. Sejna. 2008. Development and van Genuchten, M.T. 1980. A closed-form equation for predicting the hydraulic
applications of the HYDRUS and STANMOD software packages and conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 44(5):892.
related codes. Vadose Zone J. 7:587–600. doi:10.2136/vzj2007.0077 doi:10.2136/sssaj1980.03615995004400050002x
Slodicka, M. 2002. A robust and efficient linearization scheme for doubly Vereecken, H., A. Schnepf, J.W. Hopmans, M. Javaux, D. Or, T. Roose, et al. 2016.
nonlinear and degenerate parabolic problems arising in flow in porous Modeling soil processes: Review, key challenges, and new perspectives.
media. SIAM J. Sci. Comput. 23(5):1593–1614. doi:10.1137/ Vadose Zone J. 15. doi:10.2136/vzj2015.09.0131
S1064827500381860 Vogel, H.-J., and O. Ippisch. 2008. Estimation of a critical spatial discretization
Smits, K.M., A. Cihan, T. Sakaki, S.E. Howington, J.F. Peters, and T.H. limit for solving Richards’ equation at large scales. Vadose Zone J.
Illangasekare. 2013. Soil moisture and thermal behavior in the vicinity 7(1):112–114. doi:10.2136/vzj2006.0182
of buried objects affecting remote sensing detection: Experimental and Vogel, T., M.T. van Genuchten, and M. Cislerova. 2001. Effect of the shape of
modeling investigation. IEEE Trans. Geosci. Rem. Sens. 51(5):2675– the soil hydraulic functions near saturation on variably-saturated flow
2688. doi:10.1109/TGRS.2012.2214485 predictions. Adv. Water Resour. 24:133–144.
Söderlind, G. 2006. Time-step selection algorithms: Adaptivity, control, and Walker, H.F., and P. Ni. 2011. Anderson acceleration for fixed-point iterations.
signal processing. Appl. Numer. Math. 56(3-4):488–502. doi:10.1016/j. SIAM J. Numer. Anal. 49(4):1715–1735. doi:10.1137/10078356X
apnum.2005.04.026 Wang, X., and H.A. Tchelepi. 2013. Trust-region based solver for nonlinear
Söderlind, G., and L. Wang. 2006. Adaptive time-stepping and computational transport in heterogeneous porous media. J. Comput. Phys. 253:114–137.
stability. J. Comput. Appl. Math. 185(2):225–243. doi:10.1016/j. doi:10.1016/j.jcp.2013.06.041
cam.2005.03.008 Williams, G.A., C.T. Miller, and C.T. Kelley. 2000. Transformation approaches
Solin, P., and M. Kuraz. 2011. Solving the nonstationary Richards equation with for simulating flow in variably saturated porous media. Water Resour. Res.
adaptive hp-FEM. Adv. Water Resour. 34(9):1062–1081. doi:10.1016/j. 36(4):923–934. doi:10.1029/1999WR900349
advwatres.2011.04.020 Woodward, C.S., and C.N. Dawson. 2000. Analysis of expanded mixed finite
Sun, S., and M.F. Wheeler. 2006. Projections of velocity data for the compatibility element methods for a nonlinear parabolic equation modeling flow into
with transport. Comput. Methods Appl. Mech. Eng. 195(7-8):653–673. variably saturated porous media. SIAM J. Numer. Anal. 37(3):701–724.
doi:10.1016/j.cma.2005.02.011 doi:10.1137/S0036142996311040
Swartzendruber, D. 1969. The flow of water in unsaturated soils. In: R.J.M. de Yang, Z.L., G.Y. Niu, K.E. Mitchell, F. Chen, M.B. Ek, M. Barlage, and Y. Xia.
Wiest, editor, Flow through porous media. Academic Press, New York. p. 2011. The community Noah land surface model with multiparameterization
215–292. options (Noah-MP): 2. Evaluation over global river basins. J. Geophys. Res.
Talbot, C.A., and F.L. Ogden. 2008. A method for computing infiltration and Atmos. 116(D12). doi:10.1029/2010JD015140
redistribution in a discretized moisture content domain. Water Resour. Yeh, G.-T., D.-S. Shih, and J.-R.C. Cheng. 2011. An integrated media, integrated
Res. 44(8):W08453. doi:10.1029/2008WR006815 processes watershed model. Comput. Fluids 45(1):2–13. doi:10.1016/j.
Tocci, M.D., C.T. Kelley, and C.T. Miller. 1997. Accurate and economical solution compfluid.2010.11.018
of the pressure-head form of Richards’ equation by the method of lines. Younes, A., M. Fahs, and B. Belfort. 2013. Monotonicity of the cell-centred
Adv. Water Resour. 20(1):1–14. doi:10.1016/S0309-1708(96)00008-5 triangular MPFA method for saturated and unsaturated flow in
Tocci, M.D., C.T. Kelley, C.T. Miller, and C.E. Kees. 1998. Inexact heterogeneous porous media. J. Hydrol. 504:132–141. doi:10.1016/j.
Newton methods and the method of lines for solving Richards’ jhydrol.2013.09.041
equation in two space dimensions. Comput. Geosci. 2(4):291–309. Zha, Y., J. Yang, L. Yin, Y. Zhang, W. Zeng, and L. Shi. 2017. A modified
doi:10.1023/A:1011562522244 Picard iteration scheme for overcoming numerical difficulties of
Trangenstein, J.A. 2013. Numerical solution of elliptic and parabolic partial simulating infiltration into dry soil. J. Hydrol. 551:56–69. doi:10.1016/j.
differential equations with CD-ROM. Univ. Press, Cambridge. jhydrol.2017.05.053
doi:10.1017/CBO9781139025508 Zyvoloski, G. 2007. FEHM: A control volume finite element code for simulating
Trefry, M.G., and C. Muffels. 2007. Feflow: A finite-element ground water flow subsurface multi-phase multi-fluid heat and mass transfer. LAUR-07-
and transport modeling tool. Ground Water 45:525–528. doi:10.1111/ 3359. Los Alamos National Laboratory, Los Alamos, NM.
j.1745-6584.2007.00358.x

www.soils.org/publications/sssaj 1269

Das könnte Ihnen auch gefallen